0 Clusters by

May 14, 2015 - ABSTRACT: Geometric and electronic structures of linear SMnS, cyclic .... Figure 1. Three isomers for the MnS2. −/0 clusters studied ...
2 downloads 0 Views 2MB Size
Article pubs.acs.org/JPCA

Geometric and Electronic Structures for MnS2−/0 Clusters by Interpreting the Anion Photoelectron Spectrum with Quantum Chemical Calculations Van Tan Tran,† Quoc Tri Tran,† and Marc F. A. Hendrickx*,‡ †

Theoretical and Physical Chemistry Division, Dong Thap University, 783-Pham Huu Lau, Cao Lanh City, Dong Thap Vietnam Afdeling Kwantumchemie en Fysicochemie, Departement Chemie, Katholieke Universiteit Leuven, Celestijnenlaan 200F, B-3001 Heverlee-Leuven, Belgium



ABSTRACT: Geometric and electronic structures of linear SMnS, cyclic η2-MnS2, and linear η1-MnS2 isomers of MnS2− clusters have been investigated with B3LYP, CCSD(T), and NEVPT2 methods. The ground state of the anionic cluster is determined as 5Πg of the linear SMnS− isomer, while the ground state of the neutral cluster may be either the 4Σg− of the same isomer or the 6A1 of the η2-MnS2 cluster. The experimental photoelectron spectrum of the MnS2− cluster is interpreted by contributions of these two isomers. The high-intensity band at a binding energy of 2.94 eV is attributed to the 5Πg → 4Σg− transition between the linear SMnS−/0 clusters. The lower energy feature in the spectrum at binding energies between 1.9 and 2.8 eV and exhibiting a low intensity, is ascribed to electron detachments within the less stable η2-MnS2−/0 clusters. Ionizations from the lowest energy 7A1 state of this isomer to the neutral 6A1, 6A2, 8A2, and 6B2 states are responsible for this part of the spectrum. The extreme low intensity part between 1.3 and 1.9 eV can be due to excited states of either SMnS− or η2-MnS2−.



stoichiometry.13,14,17−19 For this reason reliable investigations into the nature of transition metal-containing sulfur clusters must be carried out with high-accurate quantum chemical methods. However, the most commonly applied computational methods on this class of clusters are density functional theory (DFT) techniques, chosen mainly because of their low computational cost. Moreover, for the purpose of reaching a dependable description of the multireference properties of the wave functions, CASPT2 (complete active space second-order perturbation theory) was used. On the other hand, NEVPT2 (N-electron valence state perturbation theory), which employs a different zeroth-order Hamiltonian (the Dyall Hamiltonian) than CASPT2 and is free of intruder states, has also been successfully applied to investigate transition metal-containing clusters.20−28 Further, for electronic states with a sufficient single reference character, CCSD(T) (coupled-cluster with single and double and perturbative triple excitations), with its supreme ability to recover dynamical correlation energy, has been utilized fruitfully on transition metal−sulfur clusters in order to get the correct estimation of the relative energy order of closely positioned electronic states and isomers.13,14 In this work, DFT, CCSD(T), and NEVPT2 methods will be used to investigate the complicated geometric and electronic structures of MnS2−/0 clusters in an effort to explain all features in the corresponding experimental anion photoelectron spectrum.

INTRODUCTION The geometric and electronic structures of small-sized transition metal−sulfur clusters have been extensively investigated with both experimental and computational methods, due to the fact they can serve as models to understand biological and industrial catalysis and the properties of related bulk materials.1−15 In particular, these types of clusters are usually studied by matrix-isolation infrared3−7 and photoelectron spectroscopy,8−10 while quantum chemical calculations are performed in order to understand the geometric and electronic structures by the assignments for their spectra.11−15 In this work, the MnS2−/0 clusters are investigated, for which the anion photoelectron spectrum was reported many years ago but, rather curiously, has not been understood fully yet.10,16 The spectrum, as measured at photon energy of 355 nm, shows two features with very different intensities.10 The higher energy band with the highest intensity at 2.94 eV was attributed to an electron detachment of the rigid SMnS isomer, which contains two S moieties bound to Mn.10 This assignment for the second band was later confirmed by density functional calculations.16 On the other hand, the first broad feature ranging from 1.30 to 2.80 eV with considerably lower intensity was thought to be the result of several electron detachments within the flexible η2MnS2 isomer, in which a S2 ligand binds side-on to Mn.10 This proposed assignment for the first band has to-date not been confirmed at any quantum chemical computational level. The MnS2−/0 clusters are expected to possess complicated electronic structures as is usually the case for compounds exhibiting an open 3d shell system on the transition metal center. Indeed, computations on similar type of complexes show several nearly degenerate low-lying electronic states and almost equal stabilities of different isomers for a particular © XXXX American Chemical Society

Received: March 26, 2015 Revised: May 13, 2015

A

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A



COMPUTATIONAL METHODS All three possible isomers of the studied clusters as depicted in Figure 1, were explored by calculating all low-lying electronic

ensure that the found geometries are true minima on the potential energy surface. Based on the B3LYP geometries, single point ROHF/CCSD(T) calculations were performed in order to improve the relative energies. The aug-cc-pwCVTZDK34 and aug-cc-pVTZ-DK35 basis sets are employed for Mn and S, respectively. In addition to all the valence electrons, the outer core electrons of Mn (3s and 3p electrons) were correlated. Scalar-relativistic effects were included in CCSD(T) through the second-order Douglas−Kroll Hamiltonian.36−38 All DFT calculations were performed with the ORCA package,39 while ROHF/CCSD(T) calculations were performed with the NWCHEM computational chemistry code.40 With the purpose to obtain zero-order wave functions for NEVPT2, CASSCF calculations were performed with an active space of 15 or 16 electrons in 13 orbitals. This active space was constructed by including the 3d and 4s orbitals of Mn, the 3p orbitals of S, and one virtual orbital of the total symmetric irreducible representation, in order to increase its flexibility. The aug-cc-pVTZ-DK basis set34,35 was utilized for both Mn and S. All outer core electrons of Mn (3s and 3p electrons) and the 3s of S were additionally correlated in the succeeding NEVPT2 step of the calculations. Scalar-relativistic effects were included through the second-order Douglas−Kroll Hamiltonian.36−38 As a multireference technique, NEVPT2 is particularly well suited for identifying the nature of the lowlying open shell electronic states by calculating all potential important spin multiplicities for all of the four irreducible representations of the C2v point group. Further, equilibrium structures of the lowest electronic states of the MnS2−/0 isomers were located by exploring the potential energy hypersurfaces at this multireference level. More specifically, potential energy profiles for the spectroscopically important states were constructed manually by symmetrically stretching the Mn−S

Figure 1. Three isomers for the MnS2−/0 clusters studied in this work and the corresponding coordinate systems. The Mn atom is positioned in the middle of the linear SMnS−, on the top of η2-MnS2−/0, and on the left-hand side of the η1-MnS2−/0.

states. A first isomer will be denoted as SMnS−/0 which contains two individual atomic S moieties bound to the central Mn atom (Figure 1a). The second and third isomers are the cyclic η2-MnS2−/0 (Figure 1b) and the linear η1-MnS2−/0 (Figure 1c), which are formed by a side-on and end-on binding of an S2 entity to the Mn atom, respectively. All computations for these isomers were done within the C2v point group using the coordinate systems as depicted in Figure 1. A combination of single-reference DFT, CCSD(T), and multireference NEVPT2 methods was utilized in our computational study. For DFT calculations, previous results on FeSn−/0 (n = 1−4) show that B3LYP29−31 is the better functional for sulfur-containing clusters and therefore this functional is used to study the title MnS2−/0 clusters.11−14 Unrestricted DFT were carried out with 6-311++G(3df,3pd) basis set for both Mn and S atoms.32,33 Geometries of the different low-lying electronic states of three isomers of MnS2−/0 clusters were optimized. Harmonic vibrational frequency calculations were performed to

Table 1. Geometric Parameters, Vibrational Frequencies, and Adiabatic Relative Energies (ARE) of Low-Lying Electronic States of Different Isomers of MnS2−/0 Clusters As Calculated at the B3LYP and CCSD(T) Levels method B3LYP

cluster SMnS



state 5

A2 B2 4 B1 7 A1 5 A1 6 A1 6 A2 8 A2 6 B2 7 A1 6 B1 5 A2 5 B2 4 B1 7 A1 6 A1 8 A2 6 B2 7 A1 5

SMnS η2-MnS2− η2-MnS2

CCSD(T)b

η1-MnS2− η1-MnS2 SMnS− SMnS η2-MnS2− η2-MnS2

η1-MnS2−

( Πg) (5Πg) 5

(7Σ+) (6Π)

Mn−S (Å), SMnS (deg), S−S (Å) 2.177, 2.177, 2.075, 2.350, 2.323, 2.216, 2.467, 2.489, 2.385, 2.156, 2.239,

(7Σ+)

179.13 179.00 136.70 55.25 56.27 60.29 48.24 47.80 50.41 180.00, 2.054 180.00, 1.973

frequencies (cm−1)a 89, 329, 439 44, 329, 439 56, 370, 444 263, 299, 444 242, 297, 440 293, 318, 467 270, 374, 588 216, 273, 575 254, 260, 560 296, 365(x2), 571 271, 391(x2), 616

ARE (eV)c 0.00 0.00 3.12 1.00 1.43 2.96 3.22 3.23 3.67 2.20 4.20 0.00 0.00 3.07 0.73 3.01 3.07 3.66 2.22

(0.00) (0.43) (1.96) (2.22) (2.23) (2.67)

(0.00) (2.28) (2.35) (2.93)

a

Frequencies are arranged in the order of SMnS bending, Mn−S symmetric stretching, and Mn−S asymmetric stretching modes for the linear SMnS0/− isomer; asymmetric Mn−S stretching, Mn−S symmetric stretching, and S−S stretching mode for the η2-MnS20/− isomers; and Mn−S stretching, SMnS bending, and S−S stretching mode for η1-MnS2−/0. bCCSD(T) single-point energies as calculated at the B3LYP geometries. c Values in parentheses are ARE as compared to the 7A1 of the η2-MnS2− isomer. B

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A

Table 2. Leading Configurations, Geometric Parameters, and Adiabatic and Vertical Relative Energies (ARE and VRE) of the Low-Lying States of the MnS2−/0 Clusters As Computed with NEVPT2a expt (eV) state

leading configuration

weight (%)

A2 (5Πg) 5 B2 (5Πg) 5 A1 5 B1 7 A1 7 A2 7 B1 7 B2 4 B1 (4Σg−) 4 B2 4 A2 6 B2 6 B1 6 A1 7 A1 5 A1 6 A1 6 A2 8 A2 6 B2 7 A1 (7Σ+) 6 B1 (6Π)

12a1213a1114a114b125b118b229b2210b212a223a20 12a1213a1114a114b125b118b229b2210b202a223a21 12a1113a1114a114b125b118b229b2210b212a223a21 12a1213a1114a114b115b118b229b2210b212a223a21 12a1113a1114a114b125b118b229b2210b212a223a21 12a1213a1114a114b125b118b229b2210b212a213a21 12a1213a1114a114b115b118b229b2210b212a223a21 12a1213a1114a114b125b118b229b2110b212a223a21 12a1213a1114a114b125b118b229b2210b202a223a20 12a1113a1114a114b125b118b229b2210b202a223a21 12a1113a1114a114b125b118b229b2210b212a223a20 12a1213a1114a114b115b118b229b2210b212a223a20 12a1213a1114a114b125b118b219b2210b212a213a21 12a1213a1114a114b125b118b229b2210b212a213a20 13a1114a1115a115b119b212a223a21 13a1214a1115a105b119b212a223a21 13a1114a1115a105b119b212a223a21 13a1114a1115a115b119b212a223a20 13a1114a1115a115b119b212a213a21 13a1114a1115a115b119b202a223a21 15a1116a1117a117b117b211a21 15a1116a1117a117b107b211a21

65 62 53 71 88 90 88 88 46 33 11 36 19 41 96 34 93 34 96 56 98 39

cluster SMnS−

SMnS

η2-MnS2− η2-MnS2

η1-MnS2− η1-MnS2

5

a

Mn−S (Å), SMnS (deg), S−S (Å) 2.126, 180.00 2.126, 154.50

2.073, 2.217, 2.218, 2.231, 2.242, 2.268, 2.341, 2.323, 2.209, 2.463, 2.480, 2.381, 2.141, 2.231,

140.43 180.00 180.00 180.00 180.00 180.00 55.94 57.10 62.16 48.77 48.39 51.05 180, 2.089 180, 1.991

AREa (eV) 0.00 0.09

2.67 3.32 3.34 3.44 3.51 3.58 0.96 1.77 2.89 3.03 3.04 3.44 2.58 4.34

VRE (eV)

ADE

VDE

0.00

(0.00) (0.81) (1.93) (2.07) (2.08) (2.48)

1.43 1.39 1.06 1.21 1.03 1.28 2.73 3.38 3.46 3.58 3.68 3.83 (0.00) (2.27) (2.32) (2.37) (2.62)

2.94

3.04

1.3−2.8

a

For one-electron detachment processes ARE and VRE denote the adiabatic and vertical detachment energies. Numbers in parentheses are ARE and VRE as compared to the 7A1 of the η2-MnS2− isomer.

is quite flat (vide infra). Therefore, as an acceptable approximation, the ground state of the most stable SMnS− isomer can be labeled as 5Πg. The 7A1 of the η2-MnS2− and the 7 + 7 Σ ( A1) of the η1-MnS2− are, respectively, 0.96 and 2.58 eV less stable than this anionic ground state as calculated by NEVPT2. The same relative stabilities were found with the B3LYP and CCSD(T) methods. In particular, these two septet states are placed above the anionic ground state at 1.00 and 2.20 eV by B3LYP, and at 0.73 and 2.22 eV by CCSD(T). All these results manifestly allow to deduce that the anionic ground state is the 5Πg (5A2, 5B2) state of the SMnS− isomer, which consequently is the initial state responsible for the high intensity band of the experimental photoelectron spectrum. For the neutral cluster a comparable stability for the 4B1 4 − ( Σg ) of the linear SMnS and 6A1 of the η2-MnS2 isomers was found. At the NEVPT2 level, the relative energies of these states, as compared to the anionic ground state, are 2.67 and 2.89 eV, respectively, meaning that the quartet is only 0.22 eV more stable than the sextet. However, at the single-reference B3LYP and CCSD(T) levels the relative energy order is inverted as compared to NEVPT2, with a smaller energy difference of around 0.16 eV. It should be noted that the 4B1 (4Σg−) possesses a pronounced multireference wave function with a weight for the leading configuration of 46%. Therefore, the single-reference B3LYP and CCSD(T) methods are not flexible enough to describe adequately the quartet and thereby underestimate the stability of this state. We are consequently inclined to accept that under these circumstances, NEVPT2 is more reliable than both B3LYP and CCSD(T) and, therefore, propose the 4B1 (4Σg−) state as the ground state of the neutral cluster.

bonds or bending the SMnS angle. The profiles for the bending mode allow to check the stability of the η2-MnS2−/0 isomers in respect with their lower positioned linear SMnS−/0 counterparts. The stretching mode profiles are important for the purpose of performing harmonic vibrational frequency analyses. These frequencies and equilibrium geometries were subsequently employed for Franck−Condon factor simulations. All NEVPT2 data were obtained with the DALTON suite of programs.41,42



RESULTS AND DISCUSSION Stability of Different Isomers of MnS2−/0. The relative stabilities of the various isomers, namely SMnS−/0, η2-MnS2−/0, and η1-MnS2−/0 were obtained by the B3LYP, CCSD(T), and NEVPT2 methods. The B3LYP and CCSD(T) results are collected in Table 1, while NEVPT2 results are shown in Table 2. For the anionic isomers, at all three levels of computation the most stable electronic state is identified as 5Πg (5A2, 5B2) of the linear SMnS− cluster. Other quintets states and all calculated septets, as included in Table 2, are positioned more than 1 eV higher. Most likely they are not involved in the anion photoelectron spectroscopy, so they were not further investigated at the CCSD(T) and B3LYP levels or concerning their nature. At the NEVPT2 level the 5Πg is split into the more stable quasi linear 5A2 state and the less stable bent 5B2 state with a SMnS angle of 154.50°. Apparently, the Renner−Teller effect causes the largest distortion for the latter state. The effect on the energy of such states has been known to be quite small,19 and this appears to be the case here where the splitting between the two components of 5Πg amounts to just 0.09 eV. Also the potential energy curve for the S−Mn−S bending mode C

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A Electronic Structures of MnS2−/0. The electronic structures of the MnS2−/0 clusters are most easily identified by analyzing the corresponding CASSCF wave functions. For the linear SMnS−/0 isomers, CASSCF molecular orbital plots and electron occupation numbers for the anionic 5Πg (5A2) ground state are shown in Figure 2. In this figure all

Figure 3. Pseudonatural CASSCF orbitals and the corresponding electron occupation numbers of the 7A1 state of the η2-MnS2− cluster.

Figure 2. Pseudonatural CASSCF orbitals and the corresponding electron occupation numbers for the 5A2 (5Πg) state of the SMnS− cluster.

predominantly 3p orbitals of S, namely 11a1, 12a1, 4b1, 8b2, 9b2, and 2a2, are doubly occupied. The metal 3dx2 (13a1), 3dy2−z2 (14a1), 3dxz (5b1), and 3dyz (10b2) are singly occupied, while the 3dxy (3a2) and 4s (15a1) are unoccupied. This occupation of the valence orbitals allows to schematically summarize the electronic structure of the anionic ground state formally as (S2−)(Mn3+)(S2−). By transferring an electron from 3dyz (10b2) to 3dxy (3a2), the 5B2 state can be obtained. It should be noted that these are dπ orbitals and therefore 5A2 and 5B2 are two components of a 5Πg state. Otherwise, the detachment of one electron from the 3dπ (10b2) orbital of 5Πg anionic ground state results in the 4B1 (4Σg−) neutral ground state. The CASSCF molecular orbitals of the 7A1, the lowest energy state of η2-MnS2−, are shown in Figure 3. Of the six predominantly 3p orbitals of S, five are doubly occupied (11a1, 12a1, 4b1, 8b2, and 2a2) and one is unoccupied (10b2 (σ*)). All 3d and 4s orbitals of Mn, labeled as 3dx2 (13a1), 3dy2−z2 (14a1), 4s,4p (15a1), 3dxz (5b1), 3dyz (9b2), and 3dxy (3a2) are singly occupied. This corresponds to the following formal charge distribution for this state: (Mn+)(S22−). Due to the lower oxidation state of the manganese cation, the 4s,4p orbital is occupied by one electron in this state. The lowest energy 6A1 state of η2-MnS2 can be obtained by detaching an electron from the 4s,4p (15a1) orbital of the 7A1 state of the anionic cluster. Figure 4 depicts the CASSCF molecular orbitals of the 7Σ+ 7 ( A1) for the η1-MnS2− isomer. Similar to the 7A1 of the η2MnS2− isomer, all 3d (15a1, 16a1, 7b1, 7b2, and 1a2) and 4s,4p (17a1) orbitals of Mn of 7Σ+ (7A1) are singly occupied. Concerning the six predominantly 3p orbitals of S, only the σ* (18a1) is unoccupied while the remaining orbitals (14a1, 5b1, 6b1, 5b2, and 6b2) are doubly occupied. By detaching the

Figure 4. Pseudonatural CASSCF orbitals and the corresponding electron occupation numbers of the 7A1 (7Σ+) state of the η1-MnS2− cluster.

electron from the 7b1 (dπ) orbital of the 7Σ+ state, the lowest energy 6Π (6B1) state of the neutral η1-MnS2 isomer is obtained. From the electronic structures and corresponding weights of the leading configurations of the low-lying electronic states of MnS2−/0 as included in Table 2, it is possible to deduce that the multireference character is for many relevant states quite pronounced. In particular, for the linear SMnS cluster, all states in the region of relative energies from 2.67 to 3.58 eV show a reference weight of less than 46%. The ground state 5Πg of the D

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A

Figure 5. Potential energy profiles for the symmetric Mn−S stretch and SMnS bending for SMnS−/0 (a and b, respectively), and for η2-MnS2−/0 (c and d, respectively) as obtained at the NEVPT2 level.

neutral SMnS− isomer possesses a weight for the leading configuration of around 65%. Other states in this kind are 5A1 (34%), 6A2 (34%), and 6B2 (56%) of the η2-MnS2−/0, and 6Π (6B1) (39%) of η1-MnS2. Due to this manifest multireference nature of many of the low-lying electronic states, a computational investigation of the electronic structures of MnS2−/0 requires an appropriate wave function method, such as NEVPT2. The electronic structures of the MnS2−/0 clusters are essential for an unambiguous interpretation of the anion photoelectron spectrum, in the sense that they are indispensable for identifying the one-electron detachment processes. Potential Energy Profiles of the Low-Lying States of MnS2−/0 Clusters. Starting from the B3LYP geometries, the potential energy profiles for the symmetric stretching of the two Mn−S bonds and the bending of SMnS angle were

constructed for the low-lying states of the linear SMnS−/0 and η2-MnS2−/0 clusters at the NEVPT2 level. For the SMnS−/0 isomers, the resulting potential energy profiles are displayed in Figure 5a and b, respectively. The former potential energy profile affords a bond distance of 2.126 Å for the 5Πg ground state of the anionic linear isomer. On the other hand, for the bending of SMnS angle, there is a splitting of the 5Πg due to a Renner−Teller distortion. In particular, the equilibrium geometry of the 5A2 component almost retains a linear structure, while the 5B2 component has a much smaller angle with an equilibrium value of 154.50°. However, such distortions from the linear structures have been known to be quite small in energy, as previously calculated for the 2Δg and 3Δg states of the FeO2− cluster.19 Clearly, this is also the case for the potential energy curves for the states included in Figure 5b. Most E

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A

first band, as compared to the second band, the experimental photoelectron spectrum of MnS2− was explained by a coexistence of two structural isomers. The first low intensity band was attributed to a flexible η2-MnS2 isomer, while the second band was ascribed to a rigid linear SMnS−/0 structure. Assignment of the latter band was afterward confirmed by DFT calculations with an electron affinity of 2.80 eV, which is in good agreement with the peak position of the second band.16 Rather surprisingly, as the first feature was thought to originate from the η2-MnS2−/0 clusters, the geometric and electronic structures of these clusters have not been reported yet at any level of computation. Because the 5Πg state of SMnS− is computed as the ground state of the anionic cluster, this state is used as the initial state for electron detachment processes in the photoelectron spectrum of MnS2−. From the above discussion of the electronic structures of the various electronic states, the transition from the anionic ground state 5Πg to the neutral ground state 4Σg− is accomplished by removing an electron from the 10b2 (dπ) orbital. The adiabatic detachment energy of this transition is 3.13, 3.07, and 2.67 eV as computed with B3LYP, CCSD(T), and NEVPT2, respectively. These values are in reasonable agreement with the peak position of 2.94 eV for the second band in the spectrum. The vertical detachment energy for this transition at the NEVPT2 level is 2.73 eV and therefore also corresponds well with the experimental proposed value of 3.04 eV. In general, the calculated results show that the second band in the photoelectron spectrum of MnS2− is the result of the 5Πg → 4Σg− transition within SMnS−/0 clusters. This assignment of the second band is in agreement with the previous DFT results.16 However, the nature of the responsible electronic states was not specified. In order to make sure that the second band in the spectrum is the result of the 5Πg → 4Σg− transition between the SMnS−/0 isomers, a Franck−Condon factor simulation based on the NEVPT2 potential energy profiles of Figure 5a−d, is performed. As already mentioned above, the potential energy curve for the bending mode of the 4Σg− is very flat for SMnS bond angles between 130° and 180°. Therefore, the frequencies associated with this vibrational mode will be quite small and consequently will not contribute to the shape of the band. On the other hand, the harmonic vibrational analyses of the potential energy profiles of the symmetric stretching of the linear SMnS−/0 clusters, render frequencies of 389 cm−1 for 5Πg and 382 cm−1 for 4Σg−. The Franck−Condon factor simulation for the 5Πg → 4Σg− transition, as depicted in Figure 7, shows a vibrational progression of four peaks, of which the first peak has the highest intensity. It should be noted that the silhouette of the second band in the photoelectron spectrum corresponds quite will with these results.10 Indeed, it starts with a sharp peak at 3.04 eV with a gradual decline toward higher binding energies. Overall, we can conclude that the second band in the spectrum is the result of the 5Πg → 4Σg− ionization between the linear SMnS−/0 clusters. Due to its much lower intensity than the second band, the first band with binding energy at 1.30−2.80 eV in the spectrum, should be attributed to the less stable η2-MnS2− isomer instead of the most stable linear SMnS−/0 isomer. Our NEVPT2 results in Table 2 show that the lowest energy 7A1 state of cyclic anionic cluster is 0.96 eV less stable than the 5A2 ground state of SMnS− cluster and 1.62 eV more stable than the 7Σ+ (7A1) of η1-MnS2−. Therefore, the 7A1 of the η2-MnS2− cluster is used as the initial state to explain the first band. Of course, the very low

interesting, the 5B2 possesses a flat potential energy profile, especially when the SMnS bond angle changes from 140° to 180°, with a minimum at 154.50°. However, the energy of this equilibrium geometry is only 70 cm−1 lower than the linear structure. Such a small distortion energy is even much lower than the zero point energy of 406 cm−1 of the 5B2 as calculated at the B3LYP level. Consequently, the two S atoms will vibrate around the linear structure, which implies that under all experimental conditions, a linear structure would be detected. Similar to 5B2, 4B1 also shows a very flat potential energy profile for the SMnS bending mode in the region from 130° to 180°. The energy difference between the minimum at 140.43° and the linear structure is 114 cm −1 , also meaning that experimentally a linear structure will be observed for this state. For all practical reasons, the 5B2 and 4B1 can be approximated to possess linear structures. In the remainder of this paper they are therefore referenced to as 5Πg and 4Σg−, respectively. Potential energy profiles for the η2-MnS2−/0 isomers, as constructed at the NEVPT2 level are shown in Figures 5c and d. Previous DFT results showed that there is a continuous potential energy decrease from the η2-MnS2 cluster to the linear SMnS cluster.16 However, our potential energy curves for the bending of the SMnS angle in Figure 5d, clearly show minima at 55.94° for the 7A1 state of the anionic η2-MnS2− isomer and at 62.16° for 6A1 of neutral η2-MnS2 isomer. Moreover, for the symmetric stretching of the Mn−S bonds, a minimum is located at 2.341 Å for 7A1 and of 2.209 Å for 6A1. These minima are also confirmed by B3LYP geometry optimizations and harmonic vibrational frequencies. Therefore, both the B3LYP and NEVPT2 results imply that the lowest energy states of η2MnS2−/0 isomers, namely the 7A1 and 6A1, exist as local minima on potential energy surface. By increasing the SMnS bond angles of Figure 5d further to values of 120°, it was found at the NEVPT2 level that these states have both deep potential energy wells of more than 1 eV (not depicted). This outcome is in sharp contrast to the previous DFT study that was unable to identify stable η2-MnS2 structures.16 Assignment of Photoelectron Spectrum of MnS2−. The photoelectron spectrum of MnS2− was measured with a photon energy of 355 nm or 3.49 eV. As presented in Figure 6, this spectrum shows two features.10 The first one at binding energies ranging from 1.3 to 2.8 eV is very broad and has a very low intensity. The second band has an unresolved vibrational progression due to the low resolution and exhibits a sharp peak at a binding energy of 2.94 eV. Due to the low intensity of the

Figure 6. Experimental photoelectron spectrum of MnS2− as reported in ref 10 and assignments as derived from the computational data of the present contribution. F

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A

spectroscopy for the MnS2− clusters is obviously needed to reach a more detailed picture about the structural and electronic properties, and therefore constitutes a challenge for future experimental and theoretical studies.



CONCLUSIONS The geometric and electronic structures of three isomers of MnS2−/0 clusters, namely the SMnS−/0, η2-MnS2−/0, and η1MnS2−/0, have been investigated by B3LYP, CCSD(T), and NEVPT2. The results show that for the neutral cluster, the 4Σg− state of SMnS and the 6A1 state of η2-MnS2 have almost the same stability. On the other hand, the ground state of anionic cluster is distinctly calculated as the 5Πg of the linear SMnS− isomer. Consequently, the highest intensity band at a binding energy of 2.94 eV in the experimental photoelectron spectrum of MnS2− is attributed to the one-electron detachment from the 5 Πg to the 4Σg− state within the linear SMnS isomer. The adiabatic detachment energies for this experimental feature, as calculated by DFT, CCSD(T), and NEVPT2 are respectively 3.12, 3.07, and 2.67 eV which match the experimental value of 2.94 eV very well. A Franck−Condon factor simulation for the 5 Πg → 4Σg− transition also confirms the broad shape of this band. The very low intensity feature at binding energies between 1.30 and 2.80 eV in the experimental spectrum, is partly ascribed to the electron detachments from the less stable η2-MnS2−/0 isomer. The 7A1 of η2-MnS2− is computed to 1.00, 0.73, and 0.96 eV less stable than the 5A2 ground state of the SMnS− isomer by B3LYP, CCSD(T), and NEVPT2 methods, so that this state is possibly the initial state for the electron detachments at the high-energy side of the low-intensity first band. The NEVPT2 results afford adiabatic detachment energies of 1.93, 2.07, 2.08, and 2.48 eV for transitions from 7 A1 to 6A1, 6A2, 8A2, and 6B2, respectively. Vertical detachment energies for these transitions are evaluated as 2.27, 2.32, 2.37, and 2.62 eV. These adiabatic and vertical detachment energies are in the region from 1.90 to 2.8 eV of the first band. Unfortunately, there is no one-electron detachment processes from the 7A1 of η2-MnS2− which corresponds to the starting point of the first band at binding energies at 1.30 eV. However, several transitions from excited anionic states of either SMnS− or η2-MnS2− isomers can possible be at the origin of the lowest energy part of the spectrum. Because of the experimental low intensity of this part of the spectrum no final conclusion could be reached. On the whole, the presented computational results allow to conclude that the entire photoelectron spectrum of the MnS2− cluster must be attributed to an experimental coexistence of the SMnS− and cyclic-MnS2− isomers.

Figure 7. Franck−Condon factor simulation for the 5Πg → 4Σg− transition between the ground states of the SMnS−/0 clusters at the NEVPT2 level.

intensity of the first band is then explained by the most likely low population of the 7A1 state of the η2-MnS2− cluster in the experiment. The first band is also very broad which indicates that it may exist of a few low-energy electron detachment processes for this isomer. This is indeed the case here since one-electron detachments from the 7A1 result in the 6A1, 6A2, 8 A2, and 6B2 states. The corresponding adiabatic detachment energies are 1.93, 2.07, 2.08, and 2.48 eV, while the vertical detachment energies are 2.27, 2.32, 2.37, and 2.62 eV. For these ionization processes, detachments of an electron from the 3d and 4s orbitals of Mn give rise to the three sextets, while an ionization from the predominantly 2a2 (π*) orbital of the S2 ligand results in the octet state. The lower ionization energy of 1.93 eV between the lowest states of the anionic and neutral cyclic isomers (7A1 to 6A1), when compared with the linear SMnS isomer, is the outcome of the higher oxidation of the manganese center in the latter isomers. Indeed, as a consequence, in the anionic 5Πg ground state the 4s,4p orbital is not occupied, resulting in a higher adiabatic detachment energy of 2.67 eV. In conclusion, the electron detachments processes from the 7A1 of the η2-MnS2− isomers explain the low intensity feature in the photoelectron spectrum starting from 1.90 eV to the beginning of the high intensity second band. From the discussion above it is clear that there is still no ionization identified with detachment energies of around 1.30 eV, which is the starting point of the first feature in the experimental spectrum. Because of the extremely low intensity of the spectrum in this region, it may be attributed to several ionization processes. As will become clear we cannot identify with certainty which of these transitions are indeed at the origin of the experimental signal. Just as an example, a first possible assignment corresponds to a one-electron detachment from the first electronic excited 5A1 state of the η2-MnS2− isomer. Indeed, the removal of an electron from the 13a1 (dx2) orbital of this state results in the neutral 6A1 with an adiabatic detachment energy of 1.12 eV, which agrees with the threshold of the first band of 1.30 eV. As a second possibility, a two-electron transition process from an excited state of the SMnS− isomer might also be at the origin of the low-energy part of the spectrum. It is therefore safe to conclude that due to the fact that the experimental intensity of this low-energy feature is extremely low, many electronic transitions can explain the first band in the spectrum. Also, this low intensity prevented to measure the exact starting point and to recognize possible contributions of different electron detachment processes in the region from 1.30 to 2.80 eV. Higher resolution photoelectron



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +32 16 32 73 62. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Stiefel Edward, I. Transition Metal Sulfur Chemistry: Biological and Industrial Significance and Key Trends. In Transition Metal Sulfur Chemistry; American Chemical Society, 1996; Vol. 653; pp 2−38. (2) Harris, S.; Chianelli, R. R. Catalysis by Transition Metal Sulfides: The Relation Between Calculated Electronic Trends and HDS Activity. J. Catal. 1984, 86, 400−412.

G

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry A (3) Liang, B.; Andrews, L. Infrared Spectra and Density Functional Theory Calculations of Group 4 Transition Metal Sulfides. J. Phys. Chem. A 2002, 106, 6295−6301. (4) Liang, B.; Andrews, L. Infrared Spectra and Density Functional Theory Calculations of Group 6 Transition Metal Sulfides in Solid Argon. J. Phys. Chem. A 2002, 106, 6945−6951. (5) Liang, B.; Andrews, L. Infrared Spectra and Density Functional Theory Calculations of Group V Transition Metal Sulfides. J. Phys. Chem. A 2002, 106, 3738−3743. (6) Liang, B.; Wang, X.; Andrews, L. Infrared Spectra and Density Functional Theory Calculations of Group 10 Transition Metal Sulfide Molecules and Complexes. J. Phys. Chem. A 2009, 113, 3336−3343. (7) Liang, B.; Wang, X.; Andrews, L. Infrared Spectra and Density Functional Theory Calculations of Group 8 Transition Metal Sulfide Molecules. J. Phys. Chem. A 2009, 113, 5375−5384. (8) Zhai, H.-J.; Kiran, B.; Wang, L.-S. Electronic and Structural Evolution of Monoiron Sulfur Clusters, FeSn− and FeSn (n = 1−6), from Anion Photoelectron Spectroscopy. J. Phys. Chem. A 2003, 107, 2821−2828. (9) Zhang, N.; Hayase, T.; Kawamata, H.; Nakao, K.; Nakajima, A.; Kaya, K. Photoelectron Spectroscopy of Iron-Sulfur Cluster Anions. J. Chem. Phys. 1996, 104, 3413−3419. (10) Zhang, N.; Kawamata, H.; Nakajima, A.; Kaya, K. Photoelectron Spectroscopy of Manganese-Sulfur Cluster Anions. J. Chem. Phys. 1996, 104, 36−41. (11) Clima, S.; Hendrickx, M. F. A. Interpretation of the Photoelectron Spectra of FeS2− by a Multiconfiguration Computational Approach. J. Phys. Chem. A 2007, 111, 10988−10992. (12) Clima, S.; Hendrickx, M. F. A. Photoelectron Spectra of FeS− Explained by a CASPT2 Ab Initio Study. Chem. Phys. Lett. 2007, 436, 341−345. (13) Tran, V. T.; Hendrickx, M. F. A. Assignment of the Photoelectron Spectra of FeS3− by Density Functional Theory, CASPT2, and RCCSD(T) Calculations. J. Phys. Chem. A 2011, 115, 13956−13964. (14) Tran, V. T.; Hendrickx, M. F. A. Molecular Structures for FeS4−/0 As Determined from an ab Initio Study of the Anion Photoelectron Spectra. J. Phys. Chem. A 2013, 117, 3227−3234. (15) Hübner, O.; Sauer, J. The electronic states of Fe2S2−/0/+/2+. J. Chem. Phys. 2002, 116, 617−628. (16) Dance, I. G.; Fisher, K. J. Density Functional Calculations of Electronic Structure, Geometric Structure and Stability for Molecular Manganese Sulfide Clusters. J. Chem. Soc., Dalton Trans. 1997, 2563− 2576. (17) Tran, V. T.; Hendrickx, M. F. A. Description of the Geometric and Electronic Structures Responsible for the Photoelectron Spectrum of FeO4−. J. Chem. Phys. 2011, 135, 094505. (18) Hendrickx, M. F. A.; Tran, V. T. Elucidating the Electronic Structures of the Ground States of the VO2−/0 Clusters: Synergism between Computation and Experiment. J. Chem. Theory Comput. 2014, 10, 4037−4044. (19) Hendrickx, M. F. A.; Tran, V. T. On the Electronic and Geometric Structures of FeO2−/0 and the Assignment of the Anion Photoelectron Spectrum. J. Chem. Theory Comput. 2012, 8, 3089− 3096. (20) Angeli, C.; Cimiraglia, R.; Evangelisti, S.; Leininger, T.; Malrieu, J.-P. Introduction of n-Electron Valence States for Multireference Perturbation Theory. J. Chem. Phys. 2001, 114, 10252−10264. (21) Angeli, C.; Cimiraglia, R.; Malrieu, J.-P. N-Electron Valence State Perturbation Theory: a Fast Implementation of the Strongly Contracted Variant. Chem. Phys. Lett. 2001, 350, 297−305. (22) Angeli, C.; Cimiraglia, R.; Malrieu, J.-P. N-Electron Valence State Perturbation Theory: A Spinless Formulation and an Efficient Implementation of the Strongly Contracted and of the Partially Contracted Variants. J. Chem. Phys. 2002, 117, 9138−9153. (23) Schapiro, I.; Sivalingam, K.; Neese, F. Assessment of n-Electron Valence State Perturbation Theory for Vertical Excitation Energies. J. Chem. Theory Comput. 2013, 9, 3567−3580.

(24) Angeli, C.; Cimiraglia, R. A Multireference Perturbation Theory Study on the Fe2 Molecule: in Quest of the Ground State. Mol. Phys. 2011, 109, 1503−1509. (25) Atanasov, M.; Comba, P.; Helmle, S.; Müller, D.; Neese, F. Zero-Field Splitting in a Series of Structurally Related Mononuclear NiII-Bispidine Complexes. Inorg. Chem. 2012, 51, 12324−12335. (26) Atanasov, M.; Zadrozny, J. M.; Long, J. R.; Neese, F. A Theoretical Analysis of Chemical Bonding, Vibronic Coupling, and Magnetic Anisotropy in Linear Iron(II) Complexes with SingleMolecule Magnet Behavior. Chem. Sci. 2013, 4, 139−156. (27) Camacho, C.; Witek, H. A.; Cimiraglia, R. The Low-Lying States of the Scandium Dimer. J. Chem. Phys. 2010, 132, 244306. (28) Domingo, A.; Carvajal, M.; de Graaf, C.; Sivalingam, K.; Neese, F.; Angeli, C. Metal-to-Metal Charge-Transfer Transitions: Reliable Excitation Energies from Ab Initio Calculations. Theor. Chem. Acc. 2012, 131, 1−13. (29) Becke, A. D. Density Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652. (30) Becke, A. D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098−3100. (31) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (32) McLean, A. D.; Chandler, G. S. Contracted Gaussian Basis Sets for Molecular Calculations. I. Second Row Atoms, Z=11−18. J. Chem. Phys. 1980, 72, 5639−5648. (33) Raghavachari, K.; Trucks, G. W. Highly Correlated Systems. Excitation Energies of First Row Transition Metals Sc-Cu. J. Chem. Phys. 1989, 91, 1062−1065. (34) Balabanov, N. B.; Peterson, K. A. Systematically Convergent Basis Sets for Transition Metals. I. All-Electron Correlation Consistent Basis Sets for the 3d Elements Sc-Zn. J. Chem. Phys. 2005, 123, 064107. (35) Woon, D. E.; Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. III. The Atoms Aluminum through Argon. J. Chem. Phys. 1993, 98, 1358−1371. (36) Ishikawa, Y.; Vilkas, M. J. Relativistic Quantum Mechanics of Many-electron Systems. J. Mol. Struct.: THEOCHEM 2001, 573, 139− 169. (37) Reiher, M.; Wolf, A. Exact Decoupling of the Dirac Hamiltonian. II. The Generalized Douglas-Kroll-Hess Transformation up to Arbitrary Order. J. Chem. Phys. 2004, 121, 10945−10956. (38) Reiher, M.; Wolf, A. Exact Decoupling of the Dirac Hamiltonian. I. General Theory. J. Chem. Phys. 2004, 121, 2037−2047. (39) Neese, F. The ORCA Program System. WIREs Comput. Mol. Sci. 2012, 2, 73−78. (40) Valiev, M.; Bylaska, E. J.; Govind, N.; Kowalski, K.; Straatsma, T. P.; Van Dam, H. J. J.; Wang, D.; Nieplocha, J.; Apra, E.; Windus, T. L.; et al. NWChem: A Comprehensive and Scalable Open-Source Solution for Large Scale Molecular Simulations. Comput. Phys. Commun. 2010, 181, 1477−1489. (41) Aidas, K.; Angeli, C.; Bak, K. L.; Bakken, V.; Bast, R.; Boman, L.; Christiansen, O.; Cimiraglia, R.; Coriani, S.; Dahle, P.; et al. The Dalton Quantum Chemistry Program System. WIREs Comput. Mol. Sci. 2014, 4, 269−284. (42) Dalton, a Molecular Electronic Structure Program, Release DALTON2013.0, 2013; see http://daltonprogram.org/.

H

DOI: 10.1021/acs.jpca.5b02896 J. Phys. Chem. A XXXX, XXX, XXX−XXX